You are on page 1of 22

International Journal of Numerical Methods for Heat & Fluid Flow

Numerical modeling of aeroacoustic characteristics of different savonius blade


profiles
Seyed Hamid Delbari, Amir Nejat, Mohammad H. Ahmadi, Ali Khaleghi, Marjan Goodarzi,
Article information:
To cite this document:
Seyed Hamid Delbari, Amir Nejat, Mohammad H. Ahmadi, Ali Khaleghi, Marjan Goodarzi, (2019)
"Numerical modeling of aeroacoustic characteristics of different savonius blade profiles", International
Journal of Numerical Methods for Heat & Fluid Flow, https://doi.org/10.1108/HFF-12-2018-0764
Permanent link to this document:
https://doi.org/10.1108/HFF-12-2018-0764
Downloaded by University Library At 16:44 21 June 2019 (PT)

Downloaded on: 21 June 2019, At: 16:44 (PT)


References: this document contains references to 52 other documents.
To copy this document: permissions@emeraldinsight.com
Access to this document was granted through an Emerald subscription provided by emerald-
srm:485088 []
For Authors
If you would like to write for this, or any other Emerald publication, then please use our Emerald
for Authors service information about how to choose which publication to write for and submission
guidelines are available for all. Please visit www.emeraldinsight.com/authors for more information.
About Emerald www.emeraldinsight.com
Emerald is a global publisher linking research and practice to the benefit of society. The company
manages a portfolio of more than 290 journals and over 2,350 books and book series volumes, as
well as providing an extensive range of online products and additional customer resources and
services.
Emerald is both COUNTER 4 and TRANSFER compliant. The organization is a partner of the
Committee on Publication Ethics (COPE) and also works with Portico and the LOCKSS initiative for
digital archive preservation.

*Related content and download information correct at time of download.


The current issue and full text archive of this journal is available on Emerald Insight at:
www.emeraldinsight.com/0961-5539.htm

Aeroacoustic
Numerical modeling of characteristics
aeroacoustic characteristics of
different savonius blade profiles
Seyed Hamid Delbari
Faculty of New Sciences and Technologies, University of Tehran, Tehran, Iran
Received 15 December 2018
Amir Nejat Revised 25 January 2019
Accepted 27 January 2019
School of Mechanical Engineering, University of Tehran, Tehran, Iran
Mohammad H. Ahmadi and Ali Khaleghi
Faculty of Mechanical Engineering, Shahrood University of Technology,
Downloaded by University Library At 16:44 21 June 2019 (PT)

Shahrood, Iran, and


Marjan Goodarzi
Sustainable Management of Natural Resources and Environment Research Group,
Faculty of Environment and Labour Safety, Ton Duc Thang University,
Ho Chi Minh City, Vietnam

Abstract
Purpose – This study aims to carry out numerical modeling to predict aerodynamic noise radiation from
four different Savonius rotor blade profile.
Design/methodology/approach – Incompressible unsteady reynolds-averaged navier-stokes (URANS)
approach using gamma–theta turbulence model is conducted to obtain the time accurate turbulent flow field.
The Ffowcs Williams and Hawkings (FW-H) acoustic analogy formulation is used for noise predictions at
optimal tip speed ratio (TSR).
Findings – The mean torque and power coefficients are compared with the experimental data and
acceptable agreement is observed. The total and MonoþDipole noise graphs are presented. A discrete
tonal component at low frequencies in all graphs is attributed to the blade passing frequency at the given
TSR. According to the noise prediction results, Bach type rotor has the lowest level of noise emission. The
effect of TSR on the noise level from the Bach rotor is investigated. A direct relation between angular
velocity and the noise emission is found.
Practical implications – The savonius rotor is a type of vertical axis wind turbines suited for mounting
in the vicinity of residential areas. Also, wind turbines wherein operation are efficient sources of tonal and
broadband noises and affect the inhabitable environment adversely. Therefore, the acoustic pollution
assessment is essential for the installation of wind turbines in residential areas.
Originality/value – This study aims to investigate the radiated noise level of four common Savonius
rotor blade profiles, namely, Bach type, Benesh type, semi-elliptic and conventional. As stated above,
numbers of studies exploit the URANS method coupled with the FW-H analogy to predict the
aeroacoustics behavior of wind turbines. Therefore, this approach is chosen in this research to deal with
the aeroacoustics and aerodynamic calculation of the flow field around the aforementioned Savonius blade
profiles. The effect of optimal TSR on the emitted noise and the contribution of thickness, loading and
quadrupole sources are of interest in this study.
International Journal of Numerical
Methods for Heat & Fluid Flow
Keywords CFD, Aeroacoustics, Vertical axis wind turbines, FW-H acoustic analogy, Savonius rotor © Emerald Publishing Limited
0961-5539
Paper type Research paper DOI 10.1108/HFF-12-2018-0764
HFF 1. Introduction
In recent decades, global concerns about the impact of over-exploitation of non-renewable
energy sources have increased. Pollution, global warming and diminution of the political
and economic supports regarding the usage of renewable energies is part of the
consequences of excessive consumption of conventional fuels.
Utilcentertion of renewable energies as distributed power systems is an alternative to
reduce both consumption of conventional fuels and their detrimental effects on the
environment. It also has a fair contribution to the sustainable development of societies.
Hence, investigation in the context of harvesting energy from the wind as a clean and
renewable source of energy is of paramount importance (Kamoji et al., 2009; Tartuferi et al.,
2015; Roy and Ducoin, 2016; Rolland et al., 2013; Obara et al., 2013; Saeidi et al., 2013; Danao
et al., 2013; Chong et al., 2013). Despite significant development of horizontal axis wind
turbines (HAWTs), vertical axis wind turbines (VAWTs) have not evolved to a level to be a
reliable, efficient and affordable off-grid electricity generator harvesting energy at low wind
speeds. Moreover, VAWTs mainly are installed in residential areas. Thus, considering their
Downloaded by University Library At 16:44 21 June 2019 (PT)

acoustic pollution particularly in areas with low background noise level is very important.
Consequently, there has been a focus toward the integration of small VAWTs in district
power generation systems in urban areas in recent years (Jianu et al., 2012).
The Savonius rotor was first designed in 1929 by S.J. Savonius. It delivers a number of
advantages such as inexpensive fabrication, high dynamic and static torque, omnidirectivity,
easy installation in confined spaces and suchlike (Menet, 2004; Fujisawa, 1992). However,
compared to its competitor (the Darius turbine) lacks efficient performance. Savonius rotor can
be used to improve the affordability of decentralized electricity generation while reducing the
associated environmental impacts of such systems. Based on the preceding statements, if the
low efficiency and the noise pollution of Savonius rotors could be coped with properly, it would
be a desirable choice for discrete power generation in residential areas. In the following
sections, the recent studies on aerodynamic and aeroacoustics characteristics of both HAWTs
and VAWTs especially the Savonius rotor are discussed.

1.1 Aerodynamic characteristics of savonius rotor


Vertical axis Savonius rotor is the modified shape of the Flettner cylinder, which used to propel
marine vessels (Fleming et al., 1985). For a small scale turbine, the maximum efficiency
reported by the Sandia laboratory was 24.4 per cent (Kamoji et al., 2009). Several research
studies were conducted to improve the efficiency and overall performance of the conventional
Savonius rotor.The effect of placing a deflector in front of returning bucket on the performance
of two and three-bladed Savonius rotors was investigated by Mohamed et al. (2010). The result
indicated that installing the plate at an optimal distance from the rotor, which derived from the
experiment, increases its power coefficient up to 27 per cent owing to the more stream-lined
fluid pattern flowed toward the advancing blade and lower negative torque exerted on the
returning blade. Also, two-bladed rotors proved to be far better than three-bladed rotors in
terms of performance. In another experiment, Mohamed et al. (2011) studied the effect of blade
shape profile optimization on the power coefficient. On average, 30 per cent increase of power
coefficient at all tip speed ratio (TSR) values was reported. The highest performance was
attributed to TSR of 0.7 with 40 per cent improvement of rotor power coefficient. Other
researchers have also studied the effect of deflecting plate on the Savonious rotor performance
(Kim and Gharib, 2013; Wahyudi et al., 2015; Mohamed et al., 2010).
Although, application of nozzle guide vanes and deflectors substantially improves the
Savonius rotor performance, it has an adverse effect on Omni-directivity characteristic of the
turbine. Moreover, the integration of such devices increases complexity of the fabrication
process. Given the limitations noted, recent studies have focused on modification of the Aeroacoustic
blade’s profile and reported promising results regarding the overall performance of characteristics
Savonius turbine. In an empirical research conducted by Roy and Saha (2015), performance
of four rotors each with different blade profile was investigated and an optimal profile was
proposed through parametric analysis. With the optimal design, the gain of 34.8 per cent in
maximum power coefficient over the conventional blade profile was observed. An
innovative design of Bach type Savonius rotor using four flaps and hinges to form the
bucket of the blades introduced by Tabassum and Probert (1987). The result showed 35 per
cent increase in the average static torque comparing to the basic design with solid blades.
This is because the flaps open every time the returning blade moves toward the wind
direction and reduce the drag force on the blade. Other studies were carried out on this issue
and the results showed performance improvement under both static and dynamic conditions
(Tartuferi et al., 2015; Marinic-Kragic et al., 2018).
The effects of the other operational parameters of the turbine such as aspect ratio, blade
curvature, overlapping, bucket clearance and suchlike have been studied in (Deda Altan
Downloaded by University Library At 16:44 21 June 2019 (PT)

et al., 2016; El-Askary et al., 2018; Utomo et al., 2018; Sanditya et al., 2018; Roy and Saha,
2013; Tian et al., 2018).

1.2 Wind turbine aeroacoustics


A wind turbine blade is an efficient acoustic source, which emits sounds and may cause
inconveniences for the people dwell near the wind farms. Thus, aeroacoustic characteristics
are one of the important factors that should be considered to design low-noise wind turbines.
The frequency spectrum of wind turbine noises is extended from low frequencies, which is
hardly perceivable by the human ear, to high frequencies in the normal audible range
(Kelley et al., 1985). Wind turbine noises are mostly generated via mechanical and
aerodynamic mechanisms. Major advances in the fabrication process of mechanical
equipment such as gears in recent decades, reduced the mechanical noises considerably so
that they are not prominent in the total noise level of the wind turbine. However, reduction of
the aerodynamic noises still faces crucial problems. Hence, this study is focusing on the
latter. Aerodynamic noise is composed of narrowband (tonal) and broadband components,
which are dependent on both the geometry of the turbine rotor and the blades (Paramasivam
et al., 2015).
There are five mechanisms that cause airfoil-induced noise at low Mach number fluid
flows. Convection of fluctuating vortical structures in both laminar and turbulence
boundary layers over the trailing edge of an airfoil induces broadband acoustic noises at
mid to high frequencies. Separation of the flow from the surface of the blade is responsible
for the low to relatively high-frequency broadband noises depending on whether the
separation takes place over the whole suction side surface or just near the trailing edge.
Moreover, small vortex shedding because of the developed separated region aft of a blunt
trailing edge emits tonal noise at a certain frequency, which is dependent on the shape and
thickness of the trailing edge. Finally, vortex formation near the tip area of finite lifting
wings and blades due to the pressure gradient between the suction and the pressure sides
prompts interactions between highly turbulent three-dimensional vortices and the blade
surface near the tip region, which develops an efficient high-frequency broadband noise
emitter (Brooks et al., 1989).
Generally, noise prediction has been performed using empirical or semi-empirical
relations. Because of the tremendous advancements in computing power of computers, there
is an increasing interest toward computational aeroacoustics lately. Using computational
fluid dynamics (CFD) to model wind turbines offers an accurate design tool for researchers.
HFF However, noise prediction using numerical methods still encounters serious challenges
because sound wave energy levels are often several orders of magnitude lower than that of
the flow field. To numerically solve the sound wave equation, time and grid scales must be
refined in such a way to resolve the lowest frequency of interest. As sound waves are
measured at far-field, a dense grid must be extended to the downstream of sound source to
directly resolve the transmission of sound waves. Considering the limitations stated above,
the direct approach to numerically solve the sound generation and emission phenomenon
are prohibitively expensive. The only attempt to directly visualize the sound waves in the
near field of a horizontal axis wind turbine was carried out by Arakawa et al. (2005) using
the earth simulator, which was the fastest supercomputer thitherto to solve aeroacoustic
field in a grid of about 300 million cells. Development of an inexpensive hybrid method
based on integration of Lighthill’s acoustic analogy and CFD to compute noise level is a
novel approach in aeroacoustics. In this method, unsteady fluctuating structures in the near
field of noise sources are captured with either the Reynolds averaged Navier-Stokes (RANS),
the detached eddy simulation or the large eddy simulation (LES) models while the
Downloaded by University Library At 16:44 21 June 2019 (PT)

transportation of these perturbations to the far field is taking into account by using the
Ffowcs Williams and Hawkings (FW-H) integral equation.
In recent years, researchers investigated the aeroacoustic characteristics of wind turbines
using one of the experimental, the semi-empirical and the hybrid methods. Brooks et al.
(1989) designed and conducted a series of comprehensive experiments to better understand
the nature of airfoil self-induced noise and inflow turbulence noise. They proposed semi-
empirical equations, which are a function of the flow characteristics and the airfoil geometry
to measure the sound pressure level (SPL) related to each aerodynamic noise mechanism. To
evaluate the effect of inflow turbulence conditions on the wind turbine acoustic radiation, an
extensive experiment was carried out by Buck et al. (2016) on a 2.3 MW wind turbine. The
results showed that the inflow turbulence is the main noise source at low frequencies and it
is transmitted in both upwind and downwind of the wind turbine. An increase of 6 dB in the
SPL of low-frequency noise spectrum was recorded over different turbulence intensities.
Ottermo et al. (2017) have measured the noise emitted from a 200 kW H-rotor vertical-axis
wind turbine using a spiral microphone array at four different positions, each at the distance
of about the hub height from the turbine. The noise intensity images indicated that the
rotation of the turbine blade and its supporting arms create a highly turbulent wake, which
engulfs the second blade and is the main noise source. Oerlemans et al. (2007) conducted
acoustic field measurements for a three-bladed horizontal axis wind turbine at 58 m upwind
of the wind turbine. The results showed that the outer part of the blade is a strong noise
emitter and its intensity scales up with fifth power of the local velocity. Also, the blades
become significantly noisier when their surface is tripped.
Botha et al. (2017) have proposed a new aeroacoustic prediction model for broadband
noises from a vertical axis wind turbine by incorporating both the Brook’s semi-empirical
noise formulations and CFD-based methods. The results were compared with the
experimental noise data for an existing turbine and the superiority of the proposed approach
over the empirical ones were approved. The calculated noise data showed that the inflow-
turbulence noise is a dominant source. Tadamasa and Zangeneh (2011) performed
aeroacoustics simulation of NREL Phase VI wind turbine. The result showed that loading
noise sources are most pronounced at these operational conditions and there is a direct
relation between wind speed and the noise intensity level. Study of a Darrius vertical axis
wind turbine using unsteady reynolds-averaged navier-stokes (URANS) for aerodynamic
simulation and the FW-H equation for aeroacoustic calculation performed by Mohamed
(2014). He surveyed the effect of parameters such as airfoil shape, TSR and solidity ratio on
the turbine noise spectrum. It was found that the sound intensity emitted from the turbine is Aeroacoustic
in direct relation with both the solidity ratio and the TSR. Furthermore, a novel design of characteristics
Darrius turbine with two airfoils in each blade was proposed by Mohamed (2016).
Performing a sensitivity analysis on the gap length between two airfoils at various TSRs
showed that at 60 per cent spacing the turbine has the optimal performance in respect to
aeroacoustic and aerodynamic characteristics. The numerical prediction of aerodynamic
noise radiated from an H-Darrieus vertical axis wind turbine was investigated by
Ghasemian and Nejat (2015a, 2015b), using the incompressible LES coupled with the FW-H
acoustic analogy. It was concluded that the overall sound pressure level (OASPL) varies
logarithmically with the distance from the observer. Also, it was observed that there is a
direct relation between the noise intensity and the rotational speed. More researches on
aeroacoustics of wind turbines can be found in (Wasala et al., 2015; Khalili et al., 2015;
Ghasemian and Nejat, 2015a, 2015b; Benim et al., 2018).
This paper aims to assess the aerodynamic noise characteristics of the Savonius rotor
using the hybrid model and demonstrates the benefit of implementing this approach to
Downloaded by University Library At 16:44 21 June 2019 (PT)

design low-noise VAWT rotors. As there is not any experimental acoustic measurements of
the Savonius rotor, the noise data of this study could not be compared with the actual noise
emission of the turbine. However, as we have seen in the similar studies, which applied the
hybrid method to predict the aeroacoustic noises, the accuracy of noise estimation in this
method is highly dependent to the precise resolving of the aerodynamic field around the
turbine. Accordingly, we have assumed that the aerodynamic data validation and
verification of the cases studied in this paper could fairly justify the accuracy of the
aeroacoustic results. Moreover, this paper investigates the radiated noise of four common
Savonius rotors, namely, the Bach type, the Benesh type, the semi-elliptic and the
conventional. We have used the URANS method coupled with the FW-H analogy to
calculate the aeroacoustic and the aerodynamic fields around the aforementioned Savonius
rotors. The effect of various TSR values on the emitted noise and the contribution of
thickness, loading and quadrupole sources on OASPL are of interest in this study.

2. Governing equations
Incompressible form of the mass and the momentum equations for a fluid flow in a rotating
frame of reference is applied to the simulation. (Batchelor, 1999):

@r
þ r: r !
vr¼0 (1)
@t

@ !     !
r v r þ r: r !
v r! v !
v r þ r 2! vrþ! v !
v ! r ¼ rp þ rt r þ F
@t
(2)

There are two acceleration terms represent the rotation in equation (2). One is the Coriolis
acceleration, expressed by the term 2! v ! v r and the other is the centripetal acceleration
! ! ! !
presented by the term, v  v  r . r is the radial distance from the center of rotation, ! v
is the angular velocity of the rotor domain, ! v r is the relative velocity, p is the static
!
pressure, t r is the stress tensor and F refers to external body forces.
Using RANS technic to solve the coupled set of equations (1) and (2) results in the
emergence of additional terms in the stress tensor, which are called Reynolds average
stresses. To deal with these new unknown variables and solve the Navier–Stockes equations
HFF we need another equation or set of equations that are representing turbulence stresses in the
flow field.
The turbulence model applied in this research for the closure of the coupled set of
Navier–Stockes equations and to predict turbulent effects in the transient predictions, is the
transitional k-omega shear stress transport (SST), which is described in the following
section.

2.1 Transitional shear stress transport (gamma, theta)


The correlation-based transition model combined with the SST turbulence model was first
developed by Menter et al. (2006). This model is particularly suitable when the ratio of the
laminar to the turbulent boundary layer is prominent in the flow near a solid surface, which
is the case in the Savonius rotor at high TSRs (Kacprzak et al., 2013). The model is based on
two additional transport equations, one for the intermittency ( g ) and one for the transition
onset momentum thickness Reynolds number (Reu t), coupled with the SST turbulence
model, which is comprised of transport equations for the turbulent kinetic energy (k) and the
Downloaded by University Library At 16:44 21 June 2019 (PT)

specific dissipation rate (v ). The intermittency factor g and the transition onset momentum
thickness Reynolds number Reu t are the key factors when it comes to transition modeling
(Menter et al., 2006). The intermittency is defined as the fraction of time when the flow is
turbulent. It varies between 0 and 1, which 0 corresponds to laminar flow and 1 is attributed
to turbulent flow. The point in the flow where the velocity deviates from a laminar profile is
represented as the transition onset momentum-thickness. The transport equations for the
intermittency and the momentum thickness Reynolds number are defined as:
!
    fu t
@ fu t þ @ fu t ¼ @ @ Re
r Re r Uj Re s u tðm þ m tÞ þ Pu t (3)
@t @xj @xj @xj

  
@ @   @  mt @ g
ðr g Þ þ r Uj g ¼ m þ sg þ Pg 1  Eg 1 þ Pg 2  Eg 2 (4)
@t @xj @xj @xj

Where Pg , Pu t and Eg are transition sources based on empirical correlations. m and m t are
the dynamic viscosity and turbulent viscosity, respectively.
The SST turbulence model is used for the closure of the equations (3) and (4). This is
proved to be more accurate approach for various flow regimes (i.e. adverse pressure gradient
flows, airfoils, transonic shock waves, etc.) than other one-equation or two-equation RANS-
based models.
The SST model is defined with the turbulence kinetic energy (k) and the specific
dissipation rate (v ) equations, which are given as follows (Menter et al., 2015):
!
@ @ @  m t  @k
ð r kÞ þ ð r kui Þ ¼ m þ sk þ Gk  Yk þ Sk (5)
@t @xi @xj @xj

 
@ @ @  mt  @v
ð Þ
r v þ ð r v uiÞ ¼ m þ þ Gv  Yv þ Sv (6)
@t @xi @xj s v @x
j

Respectively, G,Y,S and s in equations (5) and (6) are production rate, dissipation rate,
source term and the turbulent Prandtl number attributed to k or v quantities.
2.2 Aeroacoustics formulation Aeroacoustic
In this paper, the FW-H equation is applied for modeling the noise emission from a moving characteristics
source. Equation (7) is the mathematical representation of the FW-H equation (Williams and
Hawkings, 1969):
1 @ 2 p0 @ 2 p0 @2  @ 

 ¼ Tij H ð f Þ  Pij nj þ r ui ðun  vn Þ d ð f Þ
c0 @t
2 2 @xi 2 @xi @xj @xi
(7)
@
þ ½ r 0 vn þ r ðun  vn Þd ð f Þ
@t
Where un and ui are the fluid velocity in the normal direction of the integration surface and
in xi direction, respectively. y n and y i represent the normal velocity of the integration surface
and the surface velocity component in xi direction. H(f ) is Heaviside function and d (f ) is
Dirac delta function. Subscript 0 refers to the unagitated properties of a medium while the
difference between the real and the undisturbed properties is defined by the primed
Downloaded by University Library At 16:44 21 June 2019 (PT)

variables (e.g. p 0 = p  p0).


f = 0 indicates the boundary of a source region, which can be either a rigid body
(impermeable) or an off-surface containing the rigid body (permeable). The unit normal
vector pointing toward the exterior region (f > 0) is shown by nj, C0 is the far field sound
speed and Tij is the Lighthill’s stress tensor, given by:
Tij ¼ r ui uj þ Pij  c20 ð r  r 0 Þd ij (8)

The right-hand side of FW-H equation indicates different sound generating sources. The
first term (Lighthill’s stress tensor) shows the time-dependent momentum, viscosity and
turbulent stresses is a quadrupole noise source. Pij is the compressive stress tensor, which
includes surface viscous and aerodynamic pressure stresses:

@ui @uj 2 @uk
Pij ¼ pd ij  m þ  d ij (9)
@xj @xi 3 @xk

The second term in equation (7) is a dipole or loading noise source, caused by the external
forces. Finally, the last source term includes the effect of mass flow rate, is a monopole or
thickness noise. The complete solution of equation (7) consists of surface and volume
integrals. The contribution from monopole, dipole and partially quadrupole sources are
represented by surface integrals, while quadrupoles are indicated by volumetric integrals.
The volume integrals becomes negligible when the flow is subsonic. Thus, the acoustic
pressure p 0 in equation (7) is defined as follows:
  0   0  
p0 !x ; t ¼ pT ! x ; t þ pL ! x ;t (10)

In equation (10), t is the observer time,! x is the receiver position. The subscripts T and L. L
refer to the thickness (monopole) and loading (dipole) components, respectively and are
given as follows (di Francescantonio, 1997):
2   3
  ð"  _ # ð r U rM _ þ  2
r 0 U n þ U n_ 4 0 n c M M 5
4p p T !
0 r 0 r
x ;t ¼ 2
dS þ 3
dS
r ð 1  Mr Þ r ð1  Mr Þ
2
f ¼0 f ¼0

(11)
" # ð" #
HFF 0!  1 ð L_ r Lr  LM
4p p L x ; t ¼ 2
dS þ 2
dS
c0 rð1  Mr Þ rð1  Mr Þ
f ¼0 f ¼0
2 n  o 3
ð L rM _ þ c M  M 2
1 4 r r 0 r 5
þ 3
dS (12)
c0 r ð1  Mr Þ
2
f ¼0

where:
r
Ui ¼ vi þ ðui  vi Þ (13)
r0

^ j þ r ui ðun  vn Þ
Li ¼ Pij n (14)
Downloaded by University Library At 16:44 21 June 2019 (PT)

Subscripts r and n refer to the unit vectors in the receiver and wall-normal directions,
! ^ !
respectively. For instance: Lr ¼ L :! r ¼ Li ri and Un ¼ U :! n ¼ Ui ni . The ratio of local
surface velocity vector to the free stream sound speed is the Mach number vector Mi. It
should be noted that the integrands in equations (11) and (12) are calculated with respect to
the retarded time t , which is defined as follows:
r
t ¼t (15)
c0

Where t, r and c0 are the receiver time, distance to the receiver and the sound speed,
respectively.

3. Problem description
This study simulates the unsteady flow field and assesses the aerodynamic noise emitted
from four types of Savonius rotors. A two-dimensional incompressible unsteady
computational fluid dynamic solver, Ansys Fluent 17.2, based on finite volume is used to
solve the Navier–Stokes equations using URANS approach.
The transitional (gamma, theta) coupled with the SST model is used to model the
unsteady turbulent structures of the flow field. This model is known to be beneficial for
predictions of flow separation under adverse pressure gradient (Cebeci, 2004). Furthermore,
because of the transition of boundary layer on the rotor surface from laminar to turbulence
at the high TSRs, using fully turbulence models result in significant differences between the
experimental and the predicted torque values (Mayle, 1991).
As the free stream Mach number is low Ma = 0.018, the air can be assumed
incompressible. Thus, a pressure-based solver is used to solve the subsonic incompressible
flow field. The velocity and pressure fields are correlated using the coupled algorithm and
the gradient is discretized using the least square cell-based scheme. Furthermore, the
discretization of convective and diffusive terms are handled using second order schemes.
Bounded second order transient formulation is chosen for temporal discretization. To model
rotating blades with respect to the stationary domain a sliding mesh approach is used with a
circular interface between both the rotary and the stationary domains. The computational
domain is illustrated in Figure 1. To decrease the blockage effect and the full development of
the flow in both upstream and downstream of the rotor a dimension of 12D  6D was
assigned to the domain. A uniform and steady velocity profile were applied to the inlet (left Aeroacoustic
side) of the domain. The velocity was set as 6.2 m/s. The outlet (right side) of the domain was characteristics
modeled as a pressure outlet with zero gauge pressure. Furthermore, symmetry boundary
conditions were applied at both the top and bottom edges of the computational domain. The
symmetry boundary condition is useful as it allows the solver to consider the wall as a part
of a larger domain, avoiding the wall effects (Tian et al., 2015). The blades’ surface was
treated as a no-slip boundary condition.
The grid was generated using the Pointwise® software. The hybrid meshing technic was
used in which O-type structured combined with unstructured quadrilateral elements used
near the blade surfaces and inside the rotating zone, respectively. Stationary domain was
covered with structured quadrilateral cells. The height of first elements adjacent to the blade
surface was set in a way to satisfy the yþ < 1 criteria so the boundary layer flow is resolved
directly without using a wall function. A mesh independency study was conducted to obtain
the optimal number of cells in the domain as a trade of between the accuracy of the results
and the computational cost. Less than 1 per cent difference between the calculated torque
Downloaded by University Library At 16:44 21 June 2019 (PT)

coefficient at TSR = 0.8 was the criteria for the optimal grid selection. The grid size for the
four configurations of the rotor blades was obtained in the ranges from 110,000 to 130,000
elements. Figure 2 shows the torque coefficient values versus number of elements for the
conventional rotor in which the difference between 110,000 and 130,000 cells grid are less
than the predefined criteria so we chose the first grid for further calculations.
The numerical time step size is set to be 0.0001 s, which ensures over 97 per cent of cells
have courant number below one. A sensitivity analysis of time step on noise calculation was
performed prior to set the temporal step size and the results showed less than 0.2 dB
difference in overall sound pressure level.
Figure 3 illustrates the frequency spectrum for the semi-elliptic rotor obtained with two
different time steps for in a period of rotor revolution. As it is clear, there is a slight
discrepancy between two graphs. Although, the graph with smaller time step has more
resolution at mid and high-frequency spectrums.
Initially, calculations were run for four complete revolutions to obtain a statistically
steady state flow field. The calculations continued for one more revolution to obtain the
time-dependent surface pressure perturbations, which are used as the acoustic source data
to compute the radiated nob of the turbines. The time domain fluctuations obtained by
running the simulations were converted to the frequency domain using the fast Fourier
transform method.

Figure 1.
(a) Schematic of the
computational
domain; and (b) the
assigned boundary
conditions
HFF The experiment of Roy et al. is chosen for validation of the present research (Roy and Saha,
2015). Values of the dynamic torque for four types of rotor, namely, conventional, Bach,
Benesh and semi-elliptic are compared with the simulation results. The rotors have two
blades and simulations are performed for the TSRs of l = 0.4, 0.6, 0.8.1. The aeroacoustics
calculations are performed only for the TSR of 0.8.

4. Results
4.1 Validation
Prior to any discussion, it is of paramount importance to assess the accuracy of the
simulation results. Hence, the mean values of power and torque coefficients from Roy et al.
experiment and numerical modeling are compared with the same values obtained in the
present study. The power coefficient is the ratio of the power generated by the rotor to the
power available in the free stream and it is obtained using equation (16):
Protor Tv T Rv
CP ¼ ¼ ¼ ¼ CT  l
Downloaded by University Library At 16:44 21 June 2019 (PT)

(16)
Pavailable 12 r Su31 12 r Su21 u
Where T is the generated torque by the wind turbine, S is the area swept by the rotor, R is
the rotor radius, CT is torque coefficient and l is TSR.
Figures 4-7 depict the graphs of power and torque coefficients for each blade profile
against various TSRs. In all figures, the power coefficient has the maximum value at the

0.35

0.3

0.25
Torque Coefficient

0.2

0.15

Figure 2. 0.1
Torque coefficient
0.05
versus number of
elements for the 0
conventional rotor at 0 20,000 40,000 60,000 80,000 100,000 120,000 140,000
TSR = 0.8
Number of Cells

Figure 3.
Frequency spectrum
obtained with two
different time steps
for a complete
revolution of the
semi-elliptic rotor
TSR of 0.8. The reason is that as the angular velocity (v ) of the rotor increases beyond the Aeroacoustic
optimal value, the coanda effect is weaken and the lifting flow inside the advancing bucket characteristics
is being disturbed accordingly. So, the lifting torque decreases. Meanwhile, faster rotation of
the turbine creates a large amount of drag on the returning blade, which is a negative torque.
Consequently, the net torque reduction become more than the increasing TSR values and
considering equation (16) the power coefficient decreases (Roy, 2014).
As it is illustrated in Figures 4-7, the power and torque coefficients are in good
agreement with the experimental data of Roy et al. Although, a general overestimation
with respect to the experimental data is observed, which might be due to the higher
aerodynamic losses in the experiment caused by three-dimensional flow structures in
the vicinity of the rotor caps. Moreover, there is a noticeable difference in the torque and
power prediction, increased with the values of the TSR, between Roy’s modeling and
the present study. The reason would be the implementation of the SST k-omega
turbulence model by Roy, which does not consider the effect of the laminar to
turbulence transition in the boundary layer flow around the rotor blades and
Downloaded by University Library At 16:44 21 June 2019 (PT)

consequently results in higher torque prediction. Although, the transitional model used

0.4 CP-exp-Roy
CT-exp-Roy CP- 2D-Simul-Roy
0.8 CT- 2D-Simul-Roy CP-2D-Present Simul (γ,θ)
CT-2D-Present Simul (γ,θ) CP-2D-Present Simul (SST k-ω)
CT-2D-Present Simul (SST k-ω) 0.3
0.6
Cp
CT

0.2
0.4

0.2 0.1
Figure 4.
(a) Torque; and (b)
0
0 0.2 0.4 0.6 0.8 1 1.2
0 power coefficients of
0 0.2 0.4 0.6 0.8 1 1.2
the conventional
λ λ
Savonius rotor
(a) (b)

CP-exp-Roy
CT-exp-Roy 0.3 CP-2D-Present Simul (γ,θ)
0.6
CT-2D-Present Simul (γ,θ)

0.2
0.4
CT

Cp

0.2 0.1

Figure 5.
(a) Torque and (b)
0 0
0 0.2 0.4 0.6 0.8 1 1.2 0 0.2 0.4 0.6 0.8 1 1.2
power coefficients of
λ the semi-elliptic type
λ
Savonius rotor
(a) (b)
HFF CP-exp-Roy
CT-exp-Roy 0.4 CP- 2D-Simul-Roy
CT- 2D-Simul-Roy CP-2D-Present Simul (γ,θ)
0.8 CT-2D-Present Simul (γ,θ)

0.3
0.6

Cp
CT

0.2
0.4

0.2 0.1
Figure 6.
(a) Torque and (b)
0 0
power coefficients of 0 0.2 0.4 0.6 0.8 1 1.2 0 0.2 0.4 0.6 0.8 1 1.2
the Bach type λ λ
Savonius rotor
(a) (b)
Downloaded by University Library At 16:44 21 June 2019 (PT)

CT-exp-Roy 0.4 CP-exp-Roy


0.6
CT-2D-Present Simul (γ,θ)
CP-2D-Present Simul (γ,θ)

0.3

0.4
Cp

0.2
CT

0.2
0.1
Figure 7.
(a) Torque and (b)
power coefficients of 0 0
0 0.2 0.4 0.6 0.8 1 1.2 0 0.2 0.4 0.6 0.8 1 1.2
the Benesh type
λ λ
Savonius rotor
(a) (b)

in this study proved to be more accurate in the prediction of the aerodynamic


characteristics of the modeled rotors. To validate the latter statement, numerical
simulation of the conventional rotor was performed by applying the SST k-omega
turbulence model and the result is illustrated in Figure 4. The overestimation of power
and torque coefficients in this model is clearly visible.
Figure 8 shows the torque coefficients of the rotors versus various TSRs. The
coefficient decreases steadily as the l increases. Moreover, there is a small difference
between the Benesh and Bach type rotors except for low values of l , which the
Benesh rotor shows slightly better performance than its counterpart. Using either
Benesh or Bach type rotor would increase the average torque value around 25 per cent
at the optimal l than the conventional rotor. The conventional rotor has the lowest
torque coefficient values of all. Finally, the semi-elliptic rotor shows lower
performance than the Bach and the Benesh rotors and higher performance than the
convetional rotor.
4.2 Aeroacoustics results Aeroacoustic
Unfortunately, there is not any experimental aeroacoustic data of the Savonius rotors so far to characteristics
verify the results of aeroacoustic calculations. However, the comparison of the aerodynamic
performance results with the experimental data in the previous section proved the accuracy
and precision of the numerical focusing performed in this study. Therefore, it is possible to
conduct a qualitative study on the aeroacoustic characteristics of the selected rotors.
As stated previously, the aerodynamic noise generated by a wind turbine comprises of
monopole (thickness), dipole (loading) and quadrupole sources. The sum of monopole and dipole
noises is calculated by integrating the pressure fluctuations on the blades surface and the total
noise is computed by integrating over the interface surface. The quadrupole contribution is the
difference between the total noise and the combination of monopole and dipole noises.
Figures 9-12 shows the total and the combination of monopole and dipole noises
spectrum for each of the four Savonius blade profile. According to the graphs, there is a
significant difference at mid to high frequencies between the SPL of total noise and the
combination of monopole and dipole noises. As the Savonius rotor is mainly a drag-based
Downloaded by University Library At 16:44 21 June 2019 (PT)

device, it can be deduced – that the interaction of the turbine wake with its blades is the
main reason for SPL difference in Figures 9-12. Moreover, there is a tonal pick at the specific
frequency of 15.1 Hz, which relates to the blade-passing frequency (BPF) of the rotor. The
BPF calculated for the TSR of 0.8 is given as below:
rad 1 rad revolution
v ¼ 47:46  ’ 7:55
s 2p sec
reolution
BPF ¼ number of blade  number of
sec

BPF ¼ 2  7:55 ’ 15:1 Hz

The streamlines of flow around the Bach type rotor along with the turbulent kinetic energy
associated with the vorticities in the flow field are depicted in Figure 13. It is clearly visible
that there are several vortex zones interacting with each other and the blades
simultaneously. Formation of each vortex is related to a specific flow characteristic. For
instance, vortex shedding from the advancing blade is due to the separation of the lifting
flow from the suction side, which leads to a drop in torque value.

Bach
0.55 Benesh
Conventional
Elliptic
0.45
CT

0.35

0.25

Figure 8.
Torque coefficient
0.15
0.4 0.5 0.6 0.7 0.8 0.9 1 versus TSR for the
modeled rotors
λ
HFF

Figure 9.
Sound pressure level
of different
mechanism of noise
generation of the
Downloaded by University Library At 16:44 21 June 2019 (PT)

conventional rotor

Figure 10.
Sound pressure level
of different
mechanism of noise
generation of the
semi-elliptic rotor

Figure 11.
Sound pressure level
of different
mechanism of noise
generation of the
Bach type rotor
The OASPL for each rotor blade calculated at a receiver located 3 m downstream of the Aeroacoustic
turbines is provided in Table I. characteristics
Based on Table I, The Bach type rotor and the Benesh type rotor have the lowest and the
highest level of radiated noise, respectively. The difference of OASPL between two cases is
approximately 3 dB. Also, the difference between the total noise and the Mono þ Dipole
noise in each case is due to the contribution of quadrupole sources, which mainly originate
from the wake generated by the rotors.
According to Table I, the Benesh and semi-elliptic rotors are emitting the highest level of
noise, respectively. Figure 14 reveals that at high-frequency ranges the noise level of the
Benesh rotor is slightly higher than the semi-elliptic rotor. It is believed that the high-
frequency noise is mainly due to the presence of small turbulence fluctuations on both time
and spatial scales within the boundary layer flow.
Downloaded by University Library At 16:44 21 June 2019 (PT)

Figure 12.
Sound pressure level
of different
mechanism of noise
generation of the
Benesh type rotor

Figure 13.
(a) Vortical structures
in the wake of the
Bach type rotor and
(b) the Turbulent
Kinetic Energy of the
vortical structures
HFF Investigating the contours of velocity for the Benesh and the Bach rotors in the time span of a
complete revolution showed that duration of the lifting flow attachment to the suction side of
the advancing blade is longer for the Benesh case. Consequently, the attached boundary layer
flow interacting with the blade surface for a longer period of time and it gives rise to the high-
frequencies small-scale fluctuations. Figure 15 displays the velocity contour for both Benesh
and semi-elliptic rotor, it is evident that the lifting flow separation occurs earlier in the semi-
elliptic case. Meanwhile, as the flow is still attached to the suction side of the Benesh rotor,
the advancing blade vortex formed because of the full separation of the lifting flow, is
completely developed and detached in the semi-elliptic case.
The effect of l on the total noise spectrum of the aero-acoustically optimal rotor (Bach
type rotor) was further investigated and the results have been depicted in Figure 16. Two
main features are highlighted in the figure. First, the frequency and the intensity of the BPF
increases with the rotational speed. The frequency of pick tonal components for l = 0.4, 0.6,
0.8 and 1 are f =7.94, 11.7, 15.1 and 19.2 Hz, respectively. Second, there is a direct relation
between the OASPL and the angular velocity of the turbine.
Downloaded by University Library At 16:44 21 June 2019 (PT)

Finally, the directivity of the OASPL for Benesh, semi-elliptic and Bach type rotors have
been shown in Figure 17. Based on the figure, Benesh and Bach type rotors have the highest
and the lowest emission levels in all directions, respectively. Also, the pattern of sound
emission for all rotors is omnidirectional, which implies the highest contribution of
monopole sources to the overall noise generated than the dipole and quadrupole sources.
Moreover, the receiver located downstream of the rotor could perceive a slightly louder
sound level than the one located at the same distance upwind of the rotor due to the
convective effects.

Table I. Rotor type Mono þ Dipole (dB) Total (dB)


The OASPL of Conventional 51 59
different noise semi-elliptic 53 60
mechanisms and the Bach 52.5 58
modeled rotors Benesh 54 61

Figure 14.
Sound pressure level
of Benesh and
semi-elliptic rotors
Aeroacoustic
characteristics

Figure 15.
Velocity contours of
Downloaded by University Library At 16:44 21 June 2019 (PT)

semi-elliptic (right)
and Benesh rotors
(left), (a, c) at the
onset of separation,
(b, d) fully separation
of the flow

Figure 16.
TSR effects on the
noise generation of
the Bach type rotor

5. Conclusion
This paper studied the aerodynamic and aeroacoustics characteristics of four type of
Savonius blade rotors, namely, conventional, semi-elliptic, Bach and Benesh Type. A
transient URANS approach using the gamma-theta turbulence model for the closure of
the Navier–Stokes equations have been implemented to obtain the time-accurate
aerodynamic flow field around the rotors. The FW-H equations have been used in this
study to calculate the aerodynamically generated noise from the rotors. Simulations were
performed for 4 different TSRs, but aeroacoustics predictions were performed for the
optimal TSR of 0.8. Mean value of torque and power coefficients were verified with the
experimental data. The results proved the accuracy of the transient turbulence model to
predict the complex boundary layer flows around Savonius rotors. The graphs of total
HFF SOUND DIRECTIVITY

0
dB 61
337.5 22.5
59
315 45
57

292.5 55 67.5

53

270 51 90

Figure 17.
247.5 112.5
Sound directivity
Downloaded by University Library At 16:44 21 June 2019 (PT)

dependency of (a)
Bach, (b) Benesh and 225 135
(c) semi-elliptic type
rotors (rotor is 202.5 157.5
situated at the center Wind Direction
180
of concentric circles) → Bach Type Semi-Eliptic Benesh

and combination of Mono þ Dipole noise were provided for the four rotors and the
difference in graphs is concluded to be because of the interaction of complex structures in
the turbine’s wake with each other and the rotor blades. The tonal pick component in the
SPL was attributed to the blade passing frequency. OASPL was presented for the rotors
and showed that the Bach and Benesh rotor has the lowest and the highest level of
emitted noise, respectively. It can be observed the OASPL is 3 dB greater in the Benesh
case. The difference between SPL of Benesh and semi-elliptic rotors in the high-frequency
range were deduced to be because of the longer period of the attached flow to the suction
side of the Benesh rotor advancing blade. Effect of TSR values on the level of radiated
noise for the optimal rotor (Bach rotor) was investigated and the results showed there is a
direct relation between the angular velocity and the sound pressure intensity level. Also,
the BPF frequency shifts toward higher frequency with the increase in TSR. At last, the
sound directivity graph showed that all the rotors have the same omni-directivity
characteristic for sound emission, which is the result of the significant contribution of
monopole sources to the overall noise generated.

References
Arakawa, C., Fleig, O., Iida, M. and Shimooka, M. (2005), “Numerical approach for noise reduction of wind
turbine blade tip with earth simulator”, Journal of the Earth Simulator, Vol. 2, pp. 11-33.
Batchelor, G.K. (1999), An Introduction to Fluid Dynamics, Cambridge University Press, Cambridge.
Benim, A.C., Diederich, M., Gül, F., Oclon, P. and Taler, J. (2018), “Computational and experimental
investigation of the aerodynamics and aeroacoustics of a small wind turbine with quasi-3D
optimization”, Energy Conversion and Management, Vol. 177, pp. 143-149.
Botha, J.D.M., Shahroki, A. and Rice, H. (2017), “An implementation of an aeroacoustic prediction model
for broadband noise from a vertical axis wind turbine using a CFD informed methodology”,
Journal of Sound and Vibration, Vol. 410, pp. 389-415.
Brooks, T.F., Pope, D.S., Marcolini, M.A., Pope, S. and Marcolini, M.A. (1989), “Airfoil self-noise and Aeroacoustic
prediction”, NASA Ref. Publ. 1218, pp. 1-142.
characteristics
Buck, S., Oerlemans, S. and Palo, S. (2016), “Experimental characterization of turbulent inflow noise on
a full-scale wind turbine”, Journal of Sound and Vibration, Vol. 385, pp. 219-238.
Cebeci, T. (2004), Turbulence Models and Their Application: Efficient Numerical Methods with
Computer Programs, Horizons Pub.
Chong, W.T., Poh, S.C., Fazlizan, A., Yip, S.Y., Chang, C.K. and Hew, W.P. (2013), “Early development of
an energy recovery wind turbine generator for exhaust air system”, Applied Energy, Vol. 112,
pp. 568-575.
Danao, L.A., Eboibi, O. and Howell, R. (2013), “An experimental investigation into the influence of
unsteady wind on the performance of a vertical axis wind turbine”, Applied Energy, Vol. 107,
pp. 403-411.
Deda Altan, B., Altan, G. and Kovan, V. (2016), “Investigation of 3D printed Savonius rotor
performance”, Renewable Energy, Vol. 99, pp. 584-591.
di Francescantonio, P. (1997), “A new boundary integral formulation for the prediction of sound
Downloaded by University Library At 16:44 21 June 2019 (PT)

radiation”, Journal of Sound and Vibration, Vol. 202 No. 4, pp. 491-509.
El-Askary, W.A., Saad, A.S., AbdelSalam, A.M. and Sakr, I.M. (2018), “Investigating the performance of
a twisted modified Savonius rotor”, Journal of Wind Engineering & Industrial Aerodynamics,
Vol. 182, pp. 344-355.
Fleming, P.D., Probert, S.D. and Tanton, D. (1985), “Designs and performances of flexible and taut sail
Savonius-type wind-turbines”, Applied Energy, Vol. 19 No. 2, pp. 97-110.
Fujisawa, N. (1992), “On the torque mechanism of Savonius rotors”, Journal of Wind Engineering &
Industrial Aerodynamics, Vol. 40 No. 3, pp. 277-292.
Ghasemian, M. and Nejat, A. (2015a), “Aero-acoustics prediction of a vertical axis wind turbine using
large eddy simulation and acoustic analogy”, Energy, Vol. 88, pp. 711-717.
Ghasemian, M. and Nejat, A. (2015b), “Aerodynamic noise prediction of a horizontal axis wind turbine
using improved delayed detached eddy simulation and acoustic analogy”, Energy Convers.
Manag, Vol. 99, pp. 210-220.
Jianu, O., Rosen, M.A. and Naterer, G. (2012), “Noise pollution prevention in wind turbines: status and
recent advances”, Sustainability, Vol. 4 No. 6, pp. 1104-1117.
Kacprzak, K., Liskiewicz, G. and Sobczak, K. (2013), “Numerical investigation of conventional and
modified savonius wind turbines”, Renewable Energy, Vol. 60, pp. 578-585.
Kamoji, M.A., Kedare, S.B. and Prabhu, S.V. (2009), “Experimental investigations on single stage
modified Savonius rotor”, Applied Energy, Vol. 86 Nos 7/8, pp. 1064-1073.
Kelley, N.D., McKenna, H.E., Hemphill, R.R., Etter, C.L., Garrelts, R.C. and Linn, N.C. (1985), “Acoustic
noise associated with the MOD-1 wind turbine: its source, impact, and control”, United States
Government Publishing Office.
Khalili, F., Majumdar, P. and Zeyghami, M. (2015), “Far-field noise prediction of wind turbines at
different receivers and wind speeds: a computational study”, Volume 7B: Fluids Engineering
Systems and Technologies, p. V07BT09A051.
Kim, D. and Gharib, M. (2013), “Efficiency improvement of straight-bladed vertical-axis wind turbines
with an upstream deflector”, Journal of Wind Engineering & Industrial Aerodynamics, Vol. 115,
pp. 48-52.
Marinic-Kragic, I., Vucina, D. and Milas, Z. (2018), “Numerical workflow for 3D shape optimization and
synthesis of vertical-axis wind turbines for specified operating regimes”, Renewable Energy,
Vol. 115, pp. 113-127.
Mayle, R.E. (1991), “The 1991 IGTI scholar lecture: the role of laminar-turbulent transition in gas
turbine engines”, Journal of Turbomachinery, Vol. 113 No. 4, pp. 509.
HFF Menet, J.-L. (2004), “A double-step Savonius rotor for local production of electricity: a design study”,
Renew. Energy, Vol. 29 No. 11, pp. 1843-1862.
Menter, F.R., Smirnov, P.E., Liu, T. and Avancha, R. (2015), “A one-equation local correlation-based
transition model”, Flow, Turbulence and Combustion, Vol. 95 No. 4, pp. 583-619.
Menter, F.R., Langtry, R.B., Likki, S.R., Suzen, Y.B., Huang, P.G. and Völker, S. (2006), “A correlation-
based transition model using local variables – part I: Model formulation”, Journal of
Turbomachinery, Vol. 128 No. 3, pp. 413-422.
Mohamed, M.H. (2014), “Aero-acoustics noise evaluation of H-rotor darrieus wind turbines”, Energy,
Vol. 65, pp. 596-604.
Mohamed, M.H. (2016), “Reduction of the generated aero-acoustics noise of a vertical axis wind turbine
using CFD (computational fluid dynamics) techniques”, Energy, Vol. 96, pp. 531-544.
Mohamed, M.H., Janiga, G., Pap, E. and Thévenin, D. (2010), “Optimal performance of a modified three-
blade Savonius turbine using frontal guiding plates”, ASME Conference Proceedings, Vol. 2010
No. 44007, pp. 803-812.
Mohamed, M.H., Janiga, G., Pap, E. and Thévenin, D. (2010), “Optimization of Savonius turbines
Downloaded by University Library At 16:44 21 June 2019 (PT)

using an obstacle shielding the returning blade”, Renewable Energy, Vol. 35 No. 11,
pp. 2618-2626.
Mohamed, M.H., Janiga, G., Pap, E. and Thévenin, D. (2011), “Optimal blade shape of a modified
Savonius turbine using an obstacle shielding the returning blade”, Energy Conversion and
Management, Vol. 52 No. 1, pp. 236-242.
Obara, S., Morizane, Y. and Morel, J. (2013), “Study on method of electricity and heat storage planning
based on energy demand and tidal flow velocity forecasts for a tidal microgrid”, Applied Energy,
Vol. 111, pp. 358-373.
Oerlemans, S., Sijtsma, P. and Méndez Lopez, B. (2007), “Location and quantification of noise sources on
a wind turbine”, Journal of Sound and Vibration, Vol. 299 Nos 4/5, pp. 869-883.
Ottermo, F., Möllerström, E., Nordborg, A., Hylander, J. and Bernhoff, H. (2017), “Location of
aerodynamic noise sources from a 200 kW vertical-axis wind turbine”, Journal of Sound and
Vibration, Vol. 400, pp. 154-166.
Paramasivam, K., Rajoo, S. and Romagnoli, A. (2015), “Suppression of tonal noise in a centrifugal fan
using guide vanes”, Journal of Sound and Vibration, Vol. 357, pp. 95-106.
Rolland, S., Newton, W., Williams, A.J., Croft, T.N., Gethin, D.T. and Cross, M. (2013), “Simulations
technique for the design of a vertical axis wind turbine device with experimental validation”,
Applied Energy, Vol. 111, pp. 1195-1203.
Roy, S. (2014), “Aerodynamic performance evaluation of a novel Savonius-style wind turbine through
unsteady simulations and wind tunnel experiments”.
Roy, S. and Ducoin, A. (2016), “Unsteady analysis on the instantaneous forces and moment arms acting
on a novel Savonius-style wind turbine”, Energy Conversion and Management, Vol. 121,
pp. 281-296.
Roy, S. and Saha, U.K. (2013), “Investigations on the effect of aspect ratio into the performance of
Savonius rotors”, ASME 2013 Gas Turbine India Conference, p. V001T07A002.
Roy, S. and Saha, U.K. (2015), “Wind tunnel experiments of a newly developed two-bladed Savonius-
style wind turbine”, Applied Energy, Vol. 137, pp. 117-125.
Saeidi, D., Sedaghat, A., Alamdari, P. and Alemrajabi, A.A. (2013), “Aerodynamic design and
economical evaluation of site specific small vertical axis wind turbines”, Applied Energy,
Vol. 101, pp. 765-775.
Sanditya, T.A., Prasetyo, A., Kristiawan, B. and Hadi, S. (2018), “Effect of blade curvature angle of
Savonius horizontal axis water turbine to the power generation”, Journal of Physics: Conference
Series, Vol. 979 No. 1, pp. 12044.
Tabassum, S.A. and Probert, S.D. (1987), “Vertical-axis wind turbine: a modified design”, Applied Aeroacoustic
Energy, Vol. 28 No. 1, pp. 59-67.
characteristics
Tadamasa, A. and Zangeneh, M. (2011), “Numerical prediction of wind turbine noise”, Renewable
Energy, Vol. 36 No. 7, pp. 1902-1912.
Tartuferi, M., D’Alessandro, V., Montelpare, S. and Ricci, R. (2015), “Enhancement of Savonius wind
rotor aerodynamic performance: a computational study of new blade shapes and curtain
systems”, Energy, Vol. 79, pp. 371-384.
Tian, W., Mao, Z., Zhang, B. and Li, Y. (2018), “Shape optimization of a Savonius wind rotor with
different convex and concave sides”, Renewable Energy, Vol. 117, pp. 287-299.
Tian, W., Song, B., Vanzwieten, J.H. and Pyakurel, P. (2015), “Computational fluid dynamics prediction
of a modified Savonius wind turbine with novel blade shapes”, Energies, pp. 7915-7929.
Utomo, I.S., Tjahjana, D.D.D.P. and Hadi, S. (2018), “Experimental studies of Savonius wind turbines
with variations sizes and fin numbers towards performance”, AIP Conference Proceedings,
Vol. 30041, p. 30041.
Wahyudi, B., Soeparman, S. and Hoeijmakers, H.W.M. (2015), “Optimization design of Savonius
Downloaded by University Library At 16:44 21 June 2019 (PT)

diffuser blade with moving deflector for hydrokınetıc cross flow turbıne rotor”, Energy Procedia,
Vol. 68, pp. 244-253.
Wasala, S.H., Storey, R.C., Norris, S.E. and Cater, J.E. (2015), “Aeroacoustic noise prediction for wind
turbines using large eddy simulation”, Journal of Wind Engineering & Industrial Aerodynamics,
Vol. 145, pp. 17-29.
Williams, J.E.F. and Hawkings, D.L. (1969), “Sound generation by turbulence and surfaces in arbitrary
motion”, Philosophical Transactions of the Royal Society of London. Series A, Mathematical and
Physical Sciences, Vol. 264 No. 1151.

Corresponding author
Marjan Goodarzi can be contacted at: marjan.goodarzi@tdtu.edu.vn

For instructions on how to order reprints of this article, please visit our website:
www.emeraldgrouppublishing.com/licensing/reprints.htm
Or contact us for further details: permissions@emeraldinsight.com

You might also like